Communication pubs.acs.org/JACS

Intermolecular Zirconium-Catalyzed Hydrophosphination of Alkenes and Dienes with Primary Phosphines Michael B. Ghebreab, Christine A. Bange, and Rory Waterman* Department of Chemistry, University of Vermont, Burlington, Vermont 05405, United States S Supporting Information *

double P−H addition product (PhC2H4)2PPh (4a) as a mixture in a 3:2 ratio as measured by 31P NMR spectroscopy (eq 1).

ABSTRACT: Catalytic hydrophosphination of terminal alkenes and dienes with primary phosphines (RPH2; R = Cy, Ph) under mild conditions has been demonstrated using a zirconium complex, [κ5-N,N,N,N,C-(Me3SiNCH 2CH 2 ) 2NCH 2CH 2NSiMe2 CH]Zr (1). Exclusively anti-Markovnikov functionalized products were observed, and the catalysis is selective for either the secondary or tertiary phosphine (i.e., double hydrophosphination) products, depending on reaction conditions. The utility of the secondary phosphine products as substrates for further elaboration was demonstrated with a platinumcatalyzed asymmetric alkylation reaction.

Initial observation of the reaction by 31P NMR spectroscopy showed the formation of (N3N)ZrPHPh (2).31 Hydrophosphination products 3a and 4a were identified by 1H and 31 P NMR spectra and mass spectrometry and notably feature 31 P NMR chemical shifts of δ −52.2 and −24.1, respectively. These are anti-Markovnikov products based on 1H NMR integrations, and compound 3a displays JPH = 205 Hz in both the 1H and 31P NMR spectra, consistent with a secondary phosphine. Simple modification of this protocol allowed for a high isolated yield of either 3a or 4a. Reaction of PhPH2 with 2 equiv of styrene in the presence of 5 mol % 1 gave exclusively tertiary product 4a based on 31P NMR spectroscopy (eq 2).

D

espite the importance of organophosphines in a range of fields,1 selective formation of P−C bonds remains a synthetic challenge.2 For example, secondary phosphines are attractive compounds. These are precursors to tertiary phosphines, but these and their oxides are also useful as ligands3 and have application in materials.4 A leading metalcatalyzed route to organophosphines is hydrophopshination,5−12 and despite success in the formation of tertiary phosphines, examples of selective formation of secondary phosphines have only been reported for limited substrates.13−17 Metal-catalyzed hydrophosphination has historically been dominated by the study of late transition metals,5−12 but recent developments in early transition-metal, alkaline earth, and rare earth catalysts have shown great promise for the reaction.14,18−27 Regardless, limitations in substrate remain. For example, unactivated alkenes are essentially absent from intermolecular hydrophopshination reactions. Prior study of hydrophosphination and hydroarsination using traimidoamine-supported zirconium compounds was limited to terminal alkynes and carbodiimides.28,29 Reinvestigation of these catalysis using primary phosphines has revealed greater reactivity than other systems for the hydrophosphination of alkene substrates, including unactivated alkenes. Additionally, these catalysts have high selectivity for either the secondary or tertiary phosphine product at ambient temperature, which may enable hydrophosphination to move from a specialty synthesis reaction to a broader commercial transformation. Treatment of styrene with 1 equiv of PhPH2 at ambient temperature in the presence of 5 mol % of [κ5-N,N,N,N,C(Me3SiNCH2CH2)2NCH2CH2NSiMe2CH]Zr (1)30 in benzene-d6 solution completely consumed styrene to afford the single P−H addition product (PhC2H4)PHPh (3a) and the © XXXX American Chemical Society

Furthermore, functionalized tertiary phosphines 4b and 4c can easily be prepared from 2 equiv of substituted styrene substrate and PhPH2 in the presence of 5 mol % 1 under the same conditions as those for 4a (eq 2). More important, high selectivity for the secondary phosphine product was achieved by reacting styrene with 2 equiv of PhPH2 and 5 mol % of 1. The result was the desired secondary phosphine 3a in a 9:1 ratio with 4a. Careful distillation can return much (up to 75%) of unreacted PhPH2, which affords pure secondary phosphine free from primary and tertiary derivatives. Though relatively few d0 metal-catalyzed hydrophosphination reactions have been reported, the predominant substrate is Ph2PH.5−12 A notable exception here is Mindiola’s titanium catalyst that uses a primary phosphine substrate and appears to proceed via [2 + 2] cycloaddition to a terminal phosphinidene.14 However, this catalysis was only realized with an alkyne substrate. For these zirconium catalysts, selectivity for Received: March 28, 2014

A

dx.doi.org/10.1021/ja503036z | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society

Communication

functionalized styrenes was observed for p-methoxystyrene (Table 1, entry c), competition experiments were used to test electronic effects. Internal competition experiments favored styrene substrates bearing electron-withdrawing substituents, an observation that may suggest some degree of nucleophilic attack from the phosphide ligand. Though primary phosphines are less bulky, these zirconium catalysts require no heating and achieve similar conversions with, in some cases, lower loadings than the perhaps most related rare earth and group 2 metal catalysts.17,22−26 Because a myriad of primary phosphines are available, CyPH2 (Cy = C6H11) was also tested as a substrate with styrene (Table 1, entry i). This reaction is qualitatively slower than that of PhPH2, requiring elevated temperatures. However, CyPH2 is the largest primary phosphine for which productive chemistry has been observed with these zirconium compounds.31 Given the initial success with styrene substrates, additional substrates, including both activated and unactivated alkenes, were targeted. Unactivated and cyclic alkenes as well as dienes were tested as hydrophosphination substrates using catalyst 1 and primary phosphines (Table 2). This catalyst was active for

either the secondary or tertiary phosphine (i.e., single or double P−H addition) products by hydrophosphination under only slightly modified conditions represents a unique and convenient, atom-efficient method to obtain a variety of phosphine derivatives. This transformation was readily elaborated to a variety of substituted styrene substrates with similar selectivity for the secondary product. Treatment of the styrene substrate with 2 equiv of primary phosphines using complex 1 as the catalyst gave functionalized secondary phosphines in good to excellent conversions (Table 1) and good isolated yields in several cases (Table 1, entries a, b, h, and i). The reactions are efficient (TON = 15−19) but at modest rates (TOF = 1.25−1.58 h−1). Table 1. Zirconium-Catalyzed Intermolecular Hydrophosphination of Styrenes with Primary Phosphinesa

Table 2. Zirconium-Catalyzed Intermolecular Hydrophosphination of Alkenes and Dienes with Primary Phosphinesa

a

Conditions: 2.0 mmol of PhPH2, 5 mol % of 1, 1 mmol of diene or alkene in 1 mL of benzene-d6. bYields determined using 1H NMR after complete consumption of unsaturated substrate. Isolated yields in parentheses are reactions at 3 mmol scale. cControl reaction without 1 performed; detectable product observed after 3 d dYield is estimated due to overlapping resonances with polymeric products. eSubstrate was not completely consumed over the time noted. fCyPH2 was used.

a Reaction conditions: 2.0 mmol of PhPH2, 5 mol % of 1, 1 mmol of styrene in 1 mL of benzene-d6 at ambient temperature. bYield was calculated from 1H NMR of the crude product after complete conversion of alkene substrate. Isolated yields in parentheses are for reactions at 3 mmol scale. cCyPH2 was used and required heating to 60 °C.

terminal alkenes and dienes but failed to catalyze the hydrophosphination of internal alkenes such as 3-hexene, ciscyclooctene, or cyclohexene even at elevated temperatures and higher catalyst loadings. The hydrophosphination of 2,3-dimethyl-1,3-butadiene with PhPH2 gave the 1,4-addition product in good isolated yield with the same regioselectivity as found with calcium and titanium catalysts (Table 2, entry a).26,27 Hydrophosphination

Halogenated styrene substrates react easily regardless the position of the halogen on the phenyl ring indicating some functional tolerance (Table 1, entries e−g). Hydrophosphination of a more sterically hindered styrene such as αmethylstyrene required longer reaction times (Table 1, entry h). Because the lowest yield in the hydrophosphination of B

dx.doi.org/10.1021/ja503036z | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society

Communication

In nearly all examples, control reactions under the same conditions in the absence of 1 gave negligible product formation after 3 d. For the best substrates (e.g., styrene), up to 10% product formation can be observed after several days of heating. Hydrophosphination of methyl acrylate with PhPH2 in the absence of 1 gave complete consumption in 3 d as a mixture of products, in contrast to the catalyzed reaction. Initial demonstration of the synthetic utility of these secondary phosphine products was made through platinumcatalyzed asymmetric alkylation as reported by Glueck.35 Reaction of 4 with benzyl bromide in the presence of NaOSiMe3 and catalytic (R,R-Me-DuPhos)PtCl(Ph)36 gave the tertiary product (PhCH2CH2)PhP(CH2Ph) (6a, eq 3). The

of this substrate was also tested using CyPH2, which required longer reaction times to give the same regioisomer of the secondary phosphine product (Table 2, entry i), similar to styrene hydrophosphination (vide supra). Cyclic but more strained alkenes such as norbornene and 2,5-norbornadiene proved to be successful in hydrophosphination with PhPH2 using 1 (Table 2, entries b and c). Though the reaction with 2,5-norbornadiene underwent hydrophosphination to form functionalized secondary phosphines, a mixture of isomers was observed (Table 2, entry b). The yield for the hydrophosphination of 2,5-norbornadiene was low because of competing polymerization that hindered separation. Hydrophosphination of norbornene using 1 afforded a mixture of regioisomers, which are identical to the secondary phosphine reported in literature.32 Simple unactivated alkenes were tested and were found to be successful substrates (TON ≈ 8−10), although longer times at higher temperatures were required (Table 2, entries d, f). Ethylene was a poor substrate (Table 2, entry g), and under these conditions, competing dehydrocoupling of PhPH2 occurred.31 However, the more facile hydrophosphination of 1-hexene and ethyl vinyl ether suggests that the solubility of ethylene in benzene may also have been a factor in the poor yield. Some improvement in conversion (TON ≈ 4) over that for ethylene was observed with the more soluble though electron-deficient 3,3,3-trifluoropropene (Table 2, entry h). The hydrophosphination of unactivated alkenes is significantly less efficient than that of styrenes and dienes, but these observations represent a substantial improvement over the current art. For example, calcium and ytterbium complexes failed to catalyze hydrophosphination of unactivated alkene substrates such as 1-hexene and norbornene with Ph2PH despite success with styrene and diene substrates.26 Hydrophosphination reactions of Michael acceptors with phosphines can occur without a catalyst but suffer from lack of selectivity and require up to 30 d to reach completion.33 Conversely, hydrophosphination of methyl acrylate with PhPH2 using 1 readily formed the secondary phosphine after 6 h with high regioselectivity and isolated yield (Table 2, entry e). Mechanistic aspects of hydrophosphination have been reviewed recently, and in general, d0 metal catalysts are suggested to proceed via an insertion reaction.9 This is consistent with a prior study of 1 in the catalytic hydrophosphination of terminal alkynes and carbodiimides that suggested insertion of the unsaturated substrate into the Zr−P bond was the P−C bond forming step in the catalysis.29 A related mechanism can be tentatively proposed here for the hydrophosphination of alkenes with 1, but it is important to underscore that not all observations (e.g., norbornadiene polymerization and 1,4-addition to butadiene) are accounted for in such a proposal. Reactions performed in the presence or absence of ambient light gave identical results, which is inconsistent with a radical reaction, though the product of norbornadiene polymerization here is similar to that initiated by radicals.34 Tertiary hydrophosphination products would likely arise by the same catalytic cycle with the secondary phosphine as a substrate. Indeed, during the preparation of tertiary phosphines 4, corresponding secondary phosphines 3 are observed but then diminish over time. These observations provide a convenient hypothesis that the selectivity is driven by kinetic factors in which secondary phosphines are simply slower substrates than the primary phosphine counterparts. However, greater mechanistic study of this system is warranted.

ee of the product was scarcely measurable, but sterically encumbered phosphines are required for optimal ee values in these transformations. Thus, reaction of 5c under the same conditions gave (norbonyl)PhP(CH2Ph) (6b) in 98% yield with 61% ee as measured with a chiral reporter.37 In summary, triamidoamine-supported zirconium compounds catalyze the hydrophosphination of alkenes and dienes using primary phosphines. This catalyst can select for the tertiary or secondary phosphine products with simple modification of reaction conditions. Additionally, compound 1 shows activity for unactivated alkenes such as 1-hexene, substrates largely absent in prior reports. The ability to further elaborate on the secondary phosphine products suggests that this catalysis could have an impact on the commercial synthesis of phosphines. Further exploration of this catalysis is underway.



ASSOCIATED CONTENT

S Supporting Information *

Experimental procedures and characterization data. This material is available free of charge via the Internet at http:// pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

[email protected] Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We dedicate this paper to the memory of Gregory L. Hillhouse (1955−2014). This work was supported by the U.S. National Science Foundation through Awards CHE-0747612 and CHE1265608.



REFERENCES

(1) Corbridge, D. E. C. Phosphorus: Chemistry, Biochemistry and Technology, 6th ed.; CRC Press: Boca Raton, FL, 2013; p 1473. (2) Gladysz, J. A.; Bedford, R. B.; Fujita, M.; Gabbaï, F. P.; Goldberg, K. I.; Holland, P. L.; Kiplinger, J. L.; Krische, M. J.; Louie, J.; Lu, C. C.; Norton, J. R.; Petrukhina, M. A.; Ren, T.; Stahl, S. S.; Tilley, T. D.; Webster, C. E.; White, M. C.; Whiteker, G. T. Organometallics 2014, 33, 1505−1527.

C

dx.doi.org/10.1021/ja503036z | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Journal of the American Chemical Society

Communication

(3) Ackermann, L. Synthesis 2006, 2006, 1557−1571. (4) Rafter, E.; Gutmann, T.; Low, F.; Buntkowsky, G.; Philippot, K.; Chaudret, B.; van Leeuwen, P. W. N. M. Catal. Sci. Technol. 2013, 3, 595−599. (5) Beletskaya, I. P.; Ananikov, V. P.; Khemchyan, L. L. Catal. Met. Complexes 2011, 37, 213−264. (6) Glueck, D. S. Dalton Trans. 2008, 5276−5286. (7) Glueck, D. S. Top. Organomet. Chem. 2010, 31, 65−100. (8) Pullarkat, S. A.; Leung, P. H. Top. Organomet. Chem. 2013, 43, 145. (9) Rosenberg, L. ACS Catal. 2013, 2845−2855. (10) Tanaka, M. Top. Curr. Chem. 2004, 232, 25. (11) Waterman, R. Dalton Trans. 2009, 18−26. (12) Wicht, D. K.; Glueck, D. S. In Catalytic Heterofunctionalization; Togni, A., Grützmacher, H., Eds.; Wiley-VCH: Weinheim, 2001. (13) Gusarova, N. K.; Shaikhutdinova, S. I.; Kazantseva, T. I.; Malysheva, S. F.; Sukhov, B. G.; Belogorlova, N. A.; Dmitriev, V. I.; Trofimov, B. A. Russ. J. Gen. Chem. 2002, 72, 371−375. (14) Zhao, G.; Basuli, F.; Kilgore, U. J.; Fan, H.; Aneetha, H.; Huffman, J. C.; Wu, G.; Mindiola, D. J. J. Am. Chem. Soc. 2006, 128, 13575−13585. (15) Wicht, D. K.; Kourkine, I. V.; Kovacik, I.; Glueck, D. S.; Concolino, T. E.; Yap, G. P. A.; Incarvito, C. D.; Rheingold, A. L. Organometallics 1999, 18, 5381−5394. (16) Douglass, M. R.; Marks, T. J. J. Am. Chem. Soc. 2000, 122, 1824−1825. (17) Basalov, I. V.; Roşca, S. C.; Lyubov, D. M.; Selikhov, A. N.; Fukin, G. K.; Sarazin, Y.; Carpentier, J.-F.; Trifonov, A. A. Inorg. Chem. 2014, 53, 1654−1661. (18) Crimmin, M. R.; Barrett, A. G. M.; Hill, M. S.; Hitchcock, P. B.; Procopiou, P. A. Organometallics 2008, 27, 497−499. (19) Behrle, A. C.; Schmidt, J. A. R. Organometallics 2012, 32, 1141− 1149. (20) Zhang, W.-X.; Nishiura, M.; Mashiko, T.; Hou, Z. Chem.Eur. J. 2008, 14, 2167−2179. (21) Zhang, W.-X.; Nishiura, M.; Hou, Z. Chem. Commun. 2006, 3812−3814. (22) Crimmin, M. R.; Barrett, A. G. M.; Hill, M. S.; Hitchcock, P. B.; Procopiou, P. A. Organometallics 2007, 26, 2953−2956. (23) Liu, B.; Roisnel, T.; Carpentier, J.-F.; Sarazin, Y. Chem.Eur. J. 2013, 19, 13445−13462. (24) Liu, B.; Roisnel, T.; Carpentier, J.-F.; Sarazin, Y. Angew. Chem., Int. Ed. 2012, 51, 4943−4946. (25) Liu, B.; Carpentier, J.-F.; Sarazin, Y. Chem.Eur. J. 2012, 18, 13259−13264. (26) Hu, H.; Cui, C. Organometallics 2012, 31, 1208−1211. (27) Perrier, A.; Comte, V.; Moïse, C.; Le Gendre, P. Chem.Eur. J. 2010, 16, 64−67. (28) Roering, A. J.; Davidson, J. J.; MacMillan, S. N.; Tanski, J. M.; Waterman, R. Dalton Trans. 2008, 4488−4498. (29) Roering, A. J.; Leshinski, S. E.; Chan, S. M.; Shalumova, T.; MacMillan, S. N.; Tanski, J. M.; Waterman, R. Organometallics 2010, 29, 2557−2565. (30) Roering, A. J.; Maddox, A. F.; Elrod, L. T.; Chan, S. M.; Ghebreab, M. B.; Donovan, K. L.; Davidson, J. J.; Hughes, R. P.; Shalumova, T.; MacMillan, S. N.; Tanski, J. M.; Waterman, R. Organometallics 2008, 28, 573−581. (31) Waterman, R. Organometallics 2007, 26, 2492−2494. (32) Busacca, C. A.; Bartholomeyzik, T.; Cheekoori, S.; Raju, R.; Eriksson, M.; Kapadia, S.; Saha, A.; Zeng, X.; Senanayake, C. H. Synlett 2009, 2009, 287−291. (33) Bourumeau, K.; Gaumont, A. C.; Denis, J. M. J. Organomet. Chem. 1997, 529, 205−14. (34) Schubert, D. M.; Norman, A. D. Inorg. Chem. 1984, 23, 4130− 4131. (35) Scriban, C.; Glueck, D. S. J. Am. Chem. Soc. 2006, 128, 2788. (36) Brunker, T. J.; Blank, N. F.; Moncarz, J. R.; Scriban, C.; Anderson, B. J.; Glueck, D. S.; Zakharov, L. N.; Golen, J. A.; Sommer,

R. D.; Incarvito, C. D.; Rheingold, A. L. Organometallics 2005, 24, 2730−2746. (37) Scriban, C.; Glueck, D. S.; Zakharov, L. N.; Kassel, W. S.; DiPasquale, A. G.; Golen, J. A.; Rheingold, A. L. Organometallics 2006, 25, 5757−67.

D

dx.doi.org/10.1021/ja503036z | J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Intermolecular zirconium-catalyzed hydrophosphination of alkenes and dienes with primary phosphines.

Catalytic hydrophosphination of terminal alkenes and dienes with primary phosphines (RPH2; R = Cy, Ph) under mild conditions has been demonstrated usi...
247KB Sizes 5 Downloads 3 Views